arXiv:1805.04488v5 [math.NA] 25 Mar 2021 oyoil fteform the of ([email protected]) eiprtv eiie cuihSho fMdcn n De and Medicine of School Schulich Medicine, Perioperative htteset the that cet htare that ficients when oeee if even done adigEio.CrepnigAto:Nm fCorrespond of Name Author: Corresponding Editor. Handling (1.1) polynomi regular our write we Thus, most.” at “degree applications. h offiinso h expression the of coefficients the arxpolynomial matrix fe only often arx leri ierzto,lnaiaino matrix of linearization linearization, algebraic matrix, ubr,i hsppr yial nepeso ntemonom the in expression an Typically paper. this in numbers, n where and hr h matrices the where ahmtc,WsenUiest (lrafi[email protected]) University Western , enti ai;frLgag nepltoa ae;adf and bases; transformations. interpolational similarity common Lagrange ba for ; orthogonal Bernstein for triples standard generalized P fbt lsi n non neet e h lsi ok[ work classic the See interest. ongoing and classic both of ( z ∗ § ‡ † 1 M ujc classifications. subject AMS Abstract. e words. Key = ) nai eerhCnr o optrAgba colo Mat of School Algebra, Computer for Centre Research Ontario nai eerhCnr o optrAgba colo Mat of School Algebra, Computer for Centre Research Ontario etefrMdclEiec,Dcso nert n Clinic and Integrity Decision Evidence, Medical for Centre eevdb h dtr nMnhDyYa.Acpe o pub for Accepted Month/Day/Year. on editors the by Received nti ae,a sdn n[ in done is as paper, this In .Introduction. 1. The z P sa ievleof eigenvalue an is degree k ℓ =0 regular A det O LERI IERZTOSO ARXPOLYNOMIALS MATRIX OF LINEARIZATIONS ALGEBRAIC FOR A ( { z edefine We φ k famti polynomial matrix a of ) P z k tnadtil,rglrmti oyoil oyoilba polynomial polynomial, matrix regular triple, Standard and k UIEY .CHAN S. Y. EUNICE ( P ( m z z xrse ntemnma ai and basis the in expressed ( ) arxplnmasaecniee,ta s with is, that considered, are polynomials matrix ) z } B P by ) sntietclyzr.Mti oyoil aemn appli many have polynomials Matrix zero. identically not is H for ( k ∈ z eeaie tnadtriples standard generalized ( ) ∈ n C z P A r xrse ndffrn oyoilbss u antheore main Our bases. polynomial differing in expressed are 0 = ) n F arcswt nre rmtefield the from entries with matrices hsrpeetto a eue ncntutn so-called constructing in used be can representation This . × arxpolynomial matrix n ≤ n × z [ = 1 z A k n EEAIE TNADTRIPLES STANDARD GENERALIZED ] nodrt s h representation the use to order in , ( ≤ r qae n the and square, are 2 z 51,1A2 65D05 15A22, 65F15, P ,w osdrtecs when case the consider we ], ) B ℓ P k ℓ ( =0 om ai o oyoil fgrade of polynomials for basis a forms z ( † + ) z OETM CORLESS M. ROBERT , ) e k sdfie sflos If follows. as defined is φ C k ( ∈ z P ) C ( ntrso h eeatplnma ai.Frconvenience, For basis. polynomial relevant the of terms in n z P polynomials. × X = ) n ( , z [ rHrieitroainlbss eacutfrtepossib the for account We bases. interpolational Hermite or A e,temnma ai,adNwo nepltoa bases; interpolational Newton and basis, monomial the ses, z Y ) ] tsr,WsenUiest ([email protected]) University Western ntistry, k X ℓ degree and , rmgnrlzdsadr rpe of triples standard generalized from ∈ =0 ℓ snttezr arx hntedge of degree the then matrix, zero the not is 1 n Author ing F lIpc MDC ete,Dprmn fAetei and Anesthesia of Department Centre), (MEDICI Impact al P m k L × eaia n ttsia cecs eateto Applied of Department Sciences, Statistical and hematical φ 1 ( iaino ot/a/er adigEio:Nm of Name Editor: Handling Month/Day/Year. on lication z n k 20 P eaia n ttsia cecs etr University Western Sciences, Statistical and hematical of = ) [ ( ‡ z any z ( , n h uvy [ surveys the and ] X ] z P ) las al F AND ) z saplnma ntevariable the in polynomial a is ( , a basis ial ewl use will We . C ( sietclyzr,tedge is degree the zero, identically is z e,cmainmti,clegemti,comrade matrix, colleague matrix, companion ses, z oyoilbasis polynomial C 1 ) 1 − m EL AIESEVYERI RAFIEE LEILI sa most at is C − = 0 where , C n φ 0 k ) w iluse will (we ℓ − ( h od“rd”i hr for short is “grade” word The . z 1 sthat is m Y = ) leri linearizations algebraic L ℓ F ( 21 ( z = P = ) ain n hi td is study their and cations z φ n [ and ] salnaiaino regular a of linearization a is A k ( k P C z ( ( sgvnfor given is X ) z − h edo complex of field the , z ) 1 ) a grade has n and ( a eepesdusing expressed be can sue.W require We used. is P z o h ) the for 29 ) ( z hc od except holds which § o hoyand theory for ] B −∞ ) ( is z ) tews,if Otherwise, . z ℓ ∗ hscnbe can This . . P ℓ ihcoef- with etabulate we ). o matrix for ( z ) o the for lt of ility and , The notion of “grade” is useful even for the monomial basis, but it is especially useful if the basis is n n k n−k an interpolational basis or the Bernstein basis φk(z) = Bk (z) = k z (1 − z) , when the degree of the polynomial may not be clear from the data.  For more information about matrix polynomials, consult [27]. See also [29], [21], and consult the seminal book [20]. Linearizations using different polynomial bases were first systematically studied in [2]; for a more up-to-date treatment see [5]. Some other recent papers of interest include [4], [28], [12], [17], and [32]; this is a very active area. See also [1]. In that paper, standard triples for structured matrices are studied. In [9] several proofs are given of strict equivalence of various linearizations to the standard second linearization for the monomial basis, which we will use without much further comment in this paper.

1.1. Organization of the Paper. In Section 1.2, we establish notation, give the definitions of algebraic linearization and of generalized standard triples. We define this last in Definition 1.1 with reference to the representation in Equation (1.7). In Section 1.3, we show how to use a generalized standard triple in the construction of algebraic linearizations. We also give a proof, by construction of the necessary equivalence matrices E(z) and F (z), that algebraic linearizations are local linearizations (a very useful notion from [13]) in the same sense that its components are. In particular, if the components A and B are linearizations in the sets ΣA and ΣB respectively, then the “algebraic linearization” is a linearization in the intersection ΣA ∩ ΣB.

In Section 2, we prove our main result, giving a universal expression for generalized standard triples for the linearizations using the polynomial bases that appear in Sections 3.1, 3.2, and 3.3. In Section 3, we introduce all the polynomial bases by scalar examples. We sum up in the final section.

1.2. Notation and Definitions. Two matrix polynomials P1(z) and P2(z) are called unimodularly equivalent if there exist unimodular matrix polynomials (that is, matrix polynomials with constant nonzero ) E(z) and F (z) with P1(z) = E(z)P2(z)F (z). A matrix pencil L(z) := zC1 − C0 is called a linearization of the matrix polynomial P (z) if both C1 and C0 are of dimension N ≥ n and L(z) is unimodularly equivalent to the block diagonal matrix diag(P (z), IN−n). Two linearizations Lm(z) and Lφ(z) are called strictly equivalent if the corresponding matrices are equivalent in the following stronger sense: C1,m = EC1,φF and C0,m = E(z)C0,φF (z), with the same invertible (constant) matrices E and F . In the paper [13] we find the powerful notion of a local linearization on a subset Σ ⊂ F. In the notation of their Definition 2.1, two rational matrices G1(z) and G2(z) are said to be equivalent in a nonempty set Σ if there exist matrices R1(z) and R2(z) each nonsingular in Σ with R1(z)G1(z)R2(z) = G2(z). This notion, intended for use with matrix rational functions, is also useful for matrix polynomials, as we will see.

There is a related idea to linearization, that of “companion pencil” (A, B), where the only requirement is that det P (z) = K det (zB − A) for some nonzero constant K. A companion pencil, therefore, has the same eigenvalues as the matrix polynomial. Companion pencils that are not linearizations do not necessarily preserve eigenvectors or elementary divisors, and are less useful than linearizations.

The usual reversal2 of a matrix polynomial of grade ℓ is the polynomial rev P (z) = zℓP (z−1). A

2This definition, which is standard, is particularly appropriate for the monomial basis. The coefficients of the reversed matrix polynomial in the monomial basis are simply the same matrices in reverse order. The notion of a reversal, however, is independent of the basis used, and indeed reversals can be done differently. In [10] for instance we find a slightly different definition of reversal, appropriate for computation in a Lagrange or Hermite interpolational basis, which maps an arbitrary finite point to infinity; this difference allows for greater numerical stability. 2 linearization L(z) = zC1 − C0 of P is called a strong linearization if rev L(z) = C1 − zC0 is also a linearization of rev P (z).

N×N If L(z) = zC1 − C0 ∈ C [z] is a linearization of P (z) in z ∈ Σ, then as a necessary consequence det(P (z)) = q(z)det(L(z)) = det(zC1 − C0) for some rational q(z) whose zeros and poles lie outside Σ. The eigenvalues of P in Σ are thus computable from the generalized eigenvalues of L. For this linearized problem several standard methods are available.

A standard pair (X, T ) for a regular matrix polynomial P (z) expressed in the monomial basis with coefficients Pk is defined in [20] or in [27] as having the following properties: X has dimension n × nℓ, T has dimension nℓ × nℓ,

ℓ k (1.2) PkXT =0 , k X=0 and that the nℓ by nℓ matrix

X XT (1.3) Q =  .  .   XT ℓ−1     is nonsingular. We can then define a third matrix

0n . (1.4) Y = Q−1  .  0  n I   n   and say that the triple (X, T , Y ) is a standard triple for a monic P (z). It is pointed out in [27] that monicity of P (z) is not required for many of the formulæ to do with standard pairs (but is required for some).

Theorem 12.1.4 of [19] states that if there are matrices X, T , and Y of dimension n × nℓ, nℓ × nℓ, and nℓ × n for which

−1 −1 (1.5) P (z)= X(zIn − T ) Y

then (X, T , Y ) is a standard triple for P (z). This is also reported in Lemma 2 in [1]. There are two other representations of a matrix polynomial given a standard triple: the right , and the left canonical form. See Theorem 2.4 in [20]. However, we do not need those representations for algebraic linearization: it is the resolvent form above that we seek to generalize in this paper. The reason is that it is this formula that is used in the proof that algebraic linearizations can be performed when the component matrix polynomials are expressed in different bases.

ℓ A {φk(z)}k=0 for polynomials of grade ℓ is a set of polynomials for which there is a 3 ℓ nonsingular matrix Φ relating the polynomials φk(z) to the 1, z, ..., z . We can write this as zℓ φ (z) ℓ zℓ−1   φℓ−1(z) ℓ−2 (1.6)   = Φ z . .   .  .     .   φ0(z)       1        Frequently, we want the ℓ × ℓ matrix that only goes up to grade ℓ − 1. This should not cause confusion. This matrix forms the foundation for the proofs in [9] of strict equivalence of various linearizations. The matrix Φ is called the change-of-basis matrix and is usually exponentially ill-conditioned in the dimension. ℓ ℓ j (ℓ−j) For example, for the Bernstein polynomials φj (z)= j z (1 − z) the change-of-basis matrix has entries j ℓ ℓ ℓ−s φi,j = i / j and condition number K1 = K∞ = (ℓ + 1) s 2 where s = ⌈(ℓ − 2)/3⌉ (the notation ⌈x⌉ ℓ+1 means the ceiling of x, the least integer not smaller than x). A short computation shows K ∼ 3 ℓ/4π as ℓ → ∞ and is thus exponentially growing with the dimension. p The constructions and definitions of standard triple discussed above are apparently tied to the monomial basis because of the powers T k in Equation (1.3). We would like to relax this restriction and extend the notion of standard triple to other bases, and also to the non-monic case. In particular, we would like the following extension of Theorem 2.4 in [20] or Theorem 12.1.4 in [19] to be available: If a matrix X ∈ Cn×N , N×N N×n the linearization L(z)= zC1 − C0 ∈ C [z], and a matrix Y ∈ C satisfy

−1 −1 (1.7) P (z)= X(zC1 − C0) Y for z∈ / z(P ) (the set of polynomial eigenvalues of P ), then X, L(z), and Y form a generalized standard triple for P (z). This obviously requires regularity of P because the formula contains P −1(z).

Indeed, we simply require L(z) to be a linearization (or local linearization, restricted to some set Σ) and take this extension as a definition. Such things exist, as we will demonstrate, and are useful, as shown in [7].

A referee pointed out that a simple characterization of linearizations L(z) can be found in [17]; this can be used as a starting point, here.

Definition 1.1. Matrices X, zC1 −C0, and Y form a generalized standard triple for the regular matrix polynomial P (z) if L(z)= zC1 − C0 is a linearization of P and Equation (1.7) holds.

Note that the matrices X and Y do not depend on z, but the linearization L(z) does, albeit only linearly; we could instead have chosen to use the words “standard quadruple” to mean (X, C1, C0, Y ) where z does not appear of any of these matrices, but this quibble seems to be a matter of aesthetics only; we may use the term “triple” to refer to X, the linearization, and Y . Proposition 1.2. If L(z) is a linearization for P (z), then there exists X and Y forming a generalized standard triple with L(z) in the above sense. Proof. This is already proved in [20]. If L(z) is a linearization for P (z) then there exist unimodular −1 −1 −1 −1 matrix polynomials E(z) and F (z) with F (z)(zC1 − C0) E (z) = diag(P (z), In,..., In). By pre- T −1 multiplying by Xp = [In, 0,..., 0] and postmultiplying by Yp = [In, 0,..., 0] , we may find X = XpF (z) −1 and Y = E (z)Yp so that Equation (1.7) holds. ♮ Remark 1.3. Thus, as a referee pointed out, generalized standard triples may be read off from the proof that L(z) is indeed a linearization (we shall see this explicitly, shortly). The matrices E(z) and F (z) above 4 are not unique; given a nonsingular constant matrix B of the right dimension, diag(In, B)E(z)(zC1 − −1 C0)F (z)diag(In, B ) is also diag(P (z), In,..., In). Therefore, generalized standard triples are also not unique.

Without using the rank criterion characterization of [17], say, it is not a priori clear that if Equation (1.7) holds then L(z) is necessarily a linearization of P (z). We do have in any case, however, the following:

Lemma 1.4. If X, Y , and the matrix pencil L(z)= zC1 − C0 are such that Equation (1.7) holds, then the matrix pencil L(z) is at least a companion pencil for P (z).

−1 −1 Proof. The norm of the resolvent kP (z)k will be large if and only if k (zC1 − C0) k is large. ♮

For various reasons, we usually do not wish to invert a “leading coefficient” here; for instance, if the polynomial basis is not degree-graded, e.g. for the Bernstein basis, then in order to even look at the true leading coefficient, we have to form a particular of the existing coefficients. In floating- point arithmetic, rounding errors can disguise the rank of the resulting matrix, hence our interest in the generalization.

If X, zC1 − C0, and Y form a generalized standard triple according to our definition, then so do XU, −1 −1 V (zC1 − C0)U, and V Y for any nonsingular matrices U and V of dimension N by N. Several similarities are used very frequently. For convenience, we describe two of the most common explicitly here. Lemma 1.5 (Flipping). Put Jas the N × N “anti-identity”, also called the sip matrix, for standard 2 involutory permutation, Ji,j =0 unless i + j = N +1 when Ji,N+1−i = 1. Then J = I and the “flipped” linearization LF (z) = J(zC1 − C0)J has in its generalized standard triple the matrices XF = XJ and YF = JY . Proof. Immediate. ♮ Remark 1.6. Flipping switches both the order of the equations and the order of the variables. It obvi- ously does not change eigenvalues. Flipping, transposition, and flipping-with-transposition give four common equivalent linearizations [34].

1.3. Algebraic Linearizations . An algebraic linearization (H, DH ), as referred to in the title of this present note, is defined in [7] as a linearization (H, DH ) of a matrix polynomial h(z) = za(z)d0b(z)+ c constructed recursively from linearizations (A, DA) and (B, DB) of the lower-grade component matrix polynomials a(z) and b(z), together with constant matrices d0 and c. The paper [7] did not give an explicit unimodular pair (EH (z), FH (z)) that reduces the linearization to diag(za(z)d0b(z)+c0, IN−n), proving that the construction actually gave a linearization, so we give a method to construct them here. Without loss of generality we take d0 = In.

Theorem 1.7. If the n by n matrix polynomial a(z) has local linearization (A, DA) on ΣA with nonsin- gular pair (EA(z), FA(z)) and if the n by n matrix polynomial b(z) has local linearization (B, DB) on ΣB with nonsingular pair (EB(z), FB(z)) then the pencil zDH −H is a local linearization of h(z)= za(z)b(z)+C on ΣA ∩ ΣB, where the matrices DH and H are given as follows:

DA (1.8) DH = I  n  5 DB   and 0 A NA,n −YAcXB 0 0 (1.9) H = −XA n n,NB .  0  NB ,NA −YB B   −1 −1 T −1 Here XA = [In, 0,..., 0]FA (z) YA = EA (z)[In, 0,..., 0] and likewise XB = [In, 0,..., 0]FB (z) and −1 T YB = EA (z)[In, 0,..., 0] give the elements of the (generalized) standard triples for a(z) and b(z). Proof. We first construct

EA(z) (1.10) E (z)= I 1  n  EB (z)   and

FA(z) (1.11) F (z)= I . 1  n  FB (z)   Applying them we get

a(z) EA(z)YAcXBFB(z) I  NA−n  (1.12) E1(z)(zDH − H)F1(z)= XAFA(z) zIn .    EB(z)YB b(z)     INB −n      Simplifying and using the definitions of the matrices appearing in the standard triples, we get

a(z) c I  NA−n  (1.13) In zIn ,    In b(z)     INB −n      which is permutationally equivalent to

a(z) c I  NA−n  (1.14) b(z) In .    INB −n     In zIn      Adding to the third block column, the last block column multiplied by −b(z), we get

a(z) c I  NA−n  (1.15) 0 In ,    6 INB −n     In −zb(z) zIn      which is permutationally equivalent to

a(z) c I −zb(z) zI  n n  (1.16) INA−n .    INB −n     In      Adding to the second block row, the last block row multiplied by −z, we get

a(z) c I −zb(z)  n  (1.17) INA−n .    INB −n     In      Finally, take into account that

I −a(z) a(z) c zb(z) I za(z)b(z)+ c (1.18) n n = . I I −zb(z) I I  n  n  n   n This completes the proof. ♮

Remark 1.8. In the case that ΣA = ΣA = F, then the intersection is also F, and this establishes that “algebraic linearizations” are linearizations if their components are linearizations.

Remark 1.9. The generalized standard triple for the algebraic linearization given here has XH = T T [0, 0, XB] and YH = [YA , 0, 0] . For these linearizations, there is no notion of expressing 1 as a linear combination of anything, because this formulation is independent of particular polynomial bases, and indeed may use different bases for different submatrices.

Algebraic linearizations offer a new class of linearizations. In [7] examples are given where the eigen- value conditioning of such linearizations is exponentially better than that of the Frobenius linearization; this demonstrates that this class of linearizations potentially offers more numerically stable algorithms3 for com- puting matrix polynomial eigenvalues than standard linearizations do. The extent to which this is possible in general has not yet been explored: at this point, we only know that this can happen for some cases.

The recursive construction of algebraic linearizations relies on generalized standard triples of each of the component matrix polynomials, and (as does the unrelated paper [32]) allows different polynomial bases to be used for each component. This present note provides some explicit formulas for generalized standard triples in various bases, for reference. As one reviewer points out, these formulas could simply be obtained by reading the proofs that these linearizations are indeed linearizations; one purpose of this paper is simply convenience.

2. Expressing 1 in the basis gives the triple. If the φk(z), 0 ≤ k ≤ ℓ − 1 form a basis, we may ℓ−1 express the polynomial 1 in that basis: then 1= k=0 ekφk(z) defines the coefficients ek uniquely. Putting

(2.19) X = eℓ−1 eℓP−2 ··· e1 e0 ⊗ In

3Any algorithm that transforms a well-conditioned problem into an ill-conditioned one as a step along the way is likely to be a numerically unstable algorithm. 7 −1 −1 for an appropriate choice of basis always gives our generalized standard triple P (z)= X(zC1 − C0) Y with

T (2.20) Y = In 0n 0n ··· 0n = e1 ⊗ In .   [Here the notation e1 means the first elementary vector: although printed in bold type, it does not look very different from the scalar e1, not bold, printed earlier as a component of X.] We prove this below for all the elementary linearizations we use in this paper. Moreover, if we replace Y above with v ⊗ In for a general anszatz column vector v, then the theorem is also true for all the linearizations of [17] as well. Theorem 2.1. Let P (z) be a regular matrix polynomial. Consider a linearization L(z) of P (z) such that

(2.21) L(z) (Φℓ(z) ⊗ Im) = (e1 ⊗ Im) P (z) ,

T ℓ T where e1 = 1 0 ··· 0 ∈ C and Φℓ(z) = φℓ−1(z) ··· φ0(z) . Let x be a vector such that xΦ (z)=1 and define X = x ⊗ I and Y = e ⊗ I , then ℓ   m 1 m  (2.22) P (z)−1 = XL(z)−1Y .

Proof. The following proof, which uses an idea of an anonymous referee, is simpler than our original one. For each of the polynomial bases we examine in this paper, the linearization satisfies either

φℓ−1(z)In In φℓ−2(z)In 0n (2.23) L(z)  .  =  .  P (z) , . .      I  0   φ0(z) n   n     for degree-graded bases, or similar statements for Bernstein bases and Lagrange and Hermite interpolational bases, as follows. For the Bernstein basis, the polynomial elements in the vector on the left are multiples of ℓ−1 ℓ−1 ℓ−1 ℓ−1 T Bj (z): [ℓ/1 · Bℓ−1 (z),ℓ/2 · Bℓ−1 (z),...,ℓ/ℓ · B0 (z)] . For the Lagrange basis, the vector on the left is T 4 [w(z),ℓ0(z),ℓ1(z),...,ℓℓ(z)] . For the Hermite interpolational basis , it is the same as for the Lagrange but with the Lagrange basis elements replaced with the Hermite interpolational basis elements.

This offers an explicit slight extension of the results of [17], and their use of a general anszatz vector v generalizes our results, in the following way: They consider the set of all matrix pencils L(z) satisfying

(2.24) L(z) (Φ ⊗ In)= v ⊗ P (z)

and later characterize just which of these matrix pencils are linearizations. They did not explicitly consider Lagrange or Hermite interpolational basis polynomials, or the Bernstein basis, but as they point out their proofs go through unchanged for matrix pencils where z appears only on the diagonal of the matrix pencil, as it does for all the cases we consider here except the Bernstein case. Their use of the general anszatz vector (all the examples we have considered just use v = e1) extends our result to the case Y = v ⊗ In.

4The paper [9] did not prove that the Hermite companion pencil discussed here is in fact a linearization. We believe that it is, but there is no proof published yet. 8 Premultiplying by L−1(z) and post-multiplying by P −1(z), we have

v φℓ−1(z)In 0n φℓ−2(z)In −1 −1 (2.25) L (z)  .  =  .  P (z) . . .     0   I   n  φ0(z) n      ℓ−1 If 1= k=0 ekφk(z) is the expression of 1 in that basis, then premultiplying both sides by P X = eℓ−1In eℓ−2In ... e0In gives the theorem. Compare also Remark1.2 which gives another formula for X and Y .

Note that in the Bernstein, Lagrange, and Hermite interpolational cases, 1 can be expressed as a linear combination of the elements given; for Lagrange and Hermite the coefficient of w(z) is 0. ♮ Remark 2.2. There are linearizations not explicitly considered in this paper; for instance, a referee has pointed out that when a matrix polynomial is expressed in a basis where the elements satisfy a linear recurrence, then there is an automatic way to build what is called a CORK linearization. See [21] and [35] for details.

In what follows we examine specific cases in detail and supply specific proofs for each basis. Indeed, much of the utility of this paper is simply writing down those details, which will allow easier programming for the uses of these generalized standard triples.

3. Scalar examples of generalized standard triples. In this section, we tabulate generalized stan- dard triples for four classes of linearizations.

In the special case n = 1 and when the monomial basis is used, a linearization is usually simplified by dividing by the leading coefficient, making the result monic and the second matrix of the pair just becomes the identity. The remaining matrix is called a “companion matrix” or Frobenius companion5. Thus finding roots of a scalar polynomial can be done by finding eigenvalues of the companion matrix. Kublanovskaya calls these “accompanying pencils” in [25].

Construction of a linearization from a companion matrix is, when possible at all, a simple matter of the n×n Kronecker (tensor) product: given C1, C0 ∈ C , take C1 = C1 ⊗ In and then replace each block pkIn r×r with the corresponding matrix coefficient Pk ∈ C (the first pk, in pkIn, is the symbolic coefficient from ℓ Cr×r f ℓ p(z)= k=0 pkφk(z); the matrix coefficient Pk ∈ is from P (z)= k=0 Pkφk(z).) This will be clearer by example. P P 3.1. Bases with three-term recurrence relations. The monomial basis, the shifted monomial basis, the Taylor basis, the Newton interpolational bases, and many common orthogonal polynomial bases all have three-term recurrence relations that, except for initial cases, can be written

(3.26) zφk(z)= αkφk+1(z)+ βkφk(z)+ γkφk−1(z) .

In all cases, we have αk 6=0. For instance, the Chebyshev polynomial recurrence is usually written Tn+1(z)= 2zTn(z) − Tn−1(z) but is easily rewritten in the above form by isolating zTn(z), and all Chebyshev αk =1/2

5The Frobenius form of a matrix is related, but different: see for instance [33]. 9 φk(z) Name αk βk γk φ0 φ1 zk monomial 1 0 01 z (z − a)k shifted monomial 1 a 0 1 z − a k (z − a) /k! Taylor n +1 a 0 1 z − a k−1 j=0 (z − τj ) Newton interpolational 1 τn 0 1 z − τ0 −1 T (z) = cos k cos (z) Chebyshev 1/2 0 1/2 1 z k Q k k Pk(z) Legendre ( + 1)/(2k + 1) 0 /(2k + 1) 1 z  Table 1 A short list of three-term recurrence relations for some important polynomial bases discussed in Section 3.1. For a more comprehensive list, see The Digital Library of Mathematical Functions. These relations and others are coded in Walter Gautschi’s packages OPQ and SOPQ [18] and in the MatrixPolynomialObject implementation package in Maple (see [22]).

for k > 1. We give a selection in Table 1, and refer the reader to section 18.9 of the Digital Library of Mathematical Functions (dlmf.nist.gov) for more. See also [18].

For all such bases, we have the linearization6

p5 β4 γ4 −p4 + p5 −p3 + p5 −p2 −p1 −p0 α4 α4 α4  1   α3 β3 γ3  (3.27) zC1 − C0 = z − ,  1   α2 β2 γ2       1   α β γ     1 1 1   1   α β     0 0      (remember that αk 6=0) and (3.28) X = 00001 , T (3.29) Y = 10000 .   For instance, a Newton interpolational basis on the nodes τ0, τ1, ..., τ5 has the linearization for matrix polynomials of this grade,

P5 −P4 + τ4P5 −P3 −P2 −P1 −P0 I I τ I  n   n 3 n  (3.30) z In − In τ2In .      In   In τ1In       In  In τ0In         ℓ ℓ ℓ k ℓ−k 3.2. The Bernstein basis. The set of polynomials {Bk(z)}k=0 = k z (1 − z) is a set of ℓ + 1 polynomials each of exact degree ℓ that together forms a basis for polynomials of grade ℓ. Bernstein  polynomimals have many applications, for example in Computer Aided Geometric Design (CAGD), and many important properties including that of optimal condition number over all bases positive on [0, 1]. They do not satisfy a simple three term recurrence relation of the form discussed in Section 3.1, although they satisfy an interesting and useful “degree-elevation” recurrence, namely

n n n−1 (3.31) (j + 1)Bj+1(z) + (n − j)Bj (z)= nBj (z) ,

6For exposition, we follow Peter Lancaster’s dictum, namely that the 5 × 5 case almost always gives the idea. 10 which specifically demonstrates that a sum of Bernstein polynomials of degree n might actually have degree strictly less than n. See [14], [15], and [16] for more details of Bernstein bases.

5 5 A Bernstein companion pencil for p5(z)= k=0 pkBk(z) is P 1 −p + p −p −p −p −p 4 5 5 3 2 1 0 2 −p −p −p −p −p  1  4 3 2 1 0 4 1 0  3    (3.32) zC1 − C0 = z  1  − 1 0 ,  3     1 0   4     1     2   1 0   5     1     1    1 2 3 4 5 (3.33) X = , 5 5 5 5 5   T (3.34) Y = 10000 .   For a construction of rational E(z) and F (z) that show this is a local linearization (unless z = 1 is an eigenvalue), see [2]. The paper [30] goes further and gives a general method of constructing all strong linearizations for Bernstein matrix polynomials, including this one (which they show is strictly equivalent to the second companion form for the monomial basis, without any restrictions on z or on the leading −1 −1 coefficient). We have p (z)= X(zC1 − C0) Y if p(z) 6=0. This pencil was first analyzed in [23] and [24]. One of the present authors independently invented and implemented a version of this companion pencil in Maple (except using P T(z), and flipped from the above form) in about 2004. For a proof of numerical stability, see the original thesis [23]. The standard triple is, we believe, new to this paper.

Example 3.1 (Singular leading coefficient case).

29 8 41 41 /100 − /25 3 − /50 /100 3 (3.35) P (z)= B0 (z)+ B1 (z) 7/10 −1/100 −7/10 91/100     9 19 /10 /100 3 1 1 3 + B2 (z)+ B3 (z) . 4/5 22/25 9851/1980 0    

Expressing Equation (3.35) into the monomial basis, we have

29/100 −8/25 29/100 −8/25 849/100 −57/20 −89/20 99/50 P (z)= + z + z2 + z3 . 7/10 −1/100 7/10 −1/100 87/10 −57/20 −89/396 1/10        

Taking the determinant of the leading coefficient

−89/20 99/50 det = (−89/20) (1/10) − (99/50) (−89/396)=0 , −89/396 1/10   we can observe that leading coefficient is singular, and thus, this matrix polynomial is non-monic. The 11 standard triple for Equation (3.35) is

−9/10 −19/100 41/50 −41/100 −29/100 8/25 −17/30 43/300 41/50 −41/100 −29/100 8/25 −4/5 −22/25 7/10 −91/100 −7/10 1/100 5099/5940 −22/25 7/10 −91/100 −7/10 1/100     1 000 00 1 010 00 C0 =   C1 =    0 100 00   0 101 00       0 010 00   0 010 30       0 001 00   0 001 03         1 0  0 1   1/3 0 2/3 0 10 0 0 X = Y =   . 0 1/3 0 2/3 0 1 0 0     0 0   0 0     Then,

−1 X (zC1 − C0) Y P (z)= I2 .

3.3. The Lagrange interpolational basis. There are by now several Lagrange basis linearizations. The use of barycentric forms means that Lagrange interpolation is efficient and numerically stable and is increasing in popularity [3]. Here is the definition of the first barycentric form for interpolation of polynomials of grade ℓ on the ℓ +1 distinct nodes τk ∈ C, 0 ≤ k ≤ ℓ. Take the partial fraction decomposition of the reciprocal of the node polynomial

ℓ (3.36) w(z)= (z − τk) , k Y=0 namely

ℓ 1 β (3.37) = k w(z) z − τk k X=0 where the coefficients βk occurring in the partial fraction decomposition are called the barycentric weights. A well-known explicit formula for the βk is

ℓ −1 (3.38) βk = (τk − τj ) . j=0 jY=6 k The Lagrange basis polynomials are normally written

ℓ (3.39) ℓk(z)= βk (z − τj ) . j=0 jY=6 k

For many sets of nodes (Chebyshev nodes on [−1, 1], or roots of unity on the unit disk), the resulting interpolant is also well-conditioned, and can even be “better than optimal” [11], see also [6]. The pencil 12 we use here is “too large” and has (numerically harmless in our experience) spurious roots at infinity7; for alternative formulations, see [35], [31]. Then, the pencil is zC1 − C0 where 0 −ρ −ρ −ρ −ρ −ρ 0 0 1 2 3 4 β τ 1  0 0    β1 τ1 (3.40) zC1 − C0 = z 1 −   .   β2 τ2   1      β3 τ3   1     β τ     4 4    Rational matrices E and F demonstrating that this is indeed a local linearization, except when the nodes are eigenvalues, can be found in the appendix of [2]. More is true: by a block permutation argument, one can find constant matrices Ek and Fk with determinant ±1 such that EkL(τk)Fk = diag(Pk, IN−n), for 0 ≤ k ≤ ℓ. Interpolating these matrices polynomially (indeed, we already know the correct barycentric weights) to form polynomial matrices E(z) and F (z) we have E(z)L(z)F (z) = diag(P (z), IN−m), and moreover because the are ±1 exactly at the nodes (and therefore nonzero in certain small neighbourhoods of the nodes) we see that L(z) and P (z) are equivalent on a set Σ (not necessarily simply connected) that contains the nodes τk. Since L(z) and P (z) are also equivalent on a set that contains everything except the nodes, we see by Proposition 2.1 of [13] that L(z) is a linearization of P (z).

The X and Y for the standard triple are

(3.41) X = 011111 , T (3.42) Y = 100000 .

Notice in this case that for the linearization N = (ℓ + 2)n whiledeg p ≤ ℓ, and therefore, there are at least 2n eigenvalues at infinity. This can be inconvenient if n is at all large.

3.4. Hermite interpolational basis. The Lagrange linearization of the previous section has been extended to Hermite interpolational bases, where some of the nodes have “flowed together”, collapsing to fewer distinct nodes8.

We suppose that at each remaining distinct node τi, 0 ≤ i ≤ N −1, say, there are now si ≥ 1 consecutive ′ ′′ pieces of information known, namely P (τi), P (τi)/1!, P (τi)/2!, and so on up to the last one, the value of (si−1) the si − 1-th derivative at z = τi, namely P (τi)/(si − 1)!. The integer si is called the confluency of the node. The known pieces of information are the local Taylor coefficients of the polynomial fitting the data: f (j)(τ ) (3.43) ρ = i , 0 ≤ j ≤ s − 1 . i,j j! i

This gives 1+ ℓ = si pieces of information, determining a polynomial of grade ℓ. The barycentric weights, this time doubly indexed as β , are again computed from the partial fraction decomposition of the reciprocal P i,j 7This numerical harmlessness needs some explanation. In brief, Lagrange basis matrix polynomial eigenvalues will be well-conditioned only in a compact region determined by the interpolation nodes, and are increasingly ill-conditioned towards infinity; in practice this means only small changes in the data are needed to perturb large finite ill-conditioned eigenvalues out to infinity. Any eigenvalues produced numerically that are well outside the region determined by the interpolation nodes are likely easily perturbed all the way to infinity, and can be safely ignored. 8 A formal definition can be found in [8], for instance. The essential idea is that given two distinct pieces of data (τk,p(τk)) and (τk+1,p(τk+1)), we also know the forward difference (pk+1 − pk)/(τk+1 − τk). In the limit as one node approaches (flows towards) the other, we still know two pieces of information: p(τk) and p′(τk). Hermite interpolation captures this idea. 13 of the node polynomial

N− si−1 1 1 1 β (3.44) = = i,j . N−1 si j+1 w(z) (z − τi) i=0 (z − τi) i=0 j=0 X X For evaluation of the interpolating polynomial,Q one should use the first or second barycentric form; see [8] for details. For theoretical work with the Hermite interpolational bases, however, we can define

si−1−j −k−1 (3.45) Hi,j (z)= βi,j+kw(z)(z − τi) . k X=0 These polynomials, each of degree ℓ, form a basis (a Hermite interpolational basis, to distinguish from the Hermite orthogonal polynomials) for polynomials of grade ℓ; moreover, they generalize the Lagrange property in that only one Taylor coefficient at only one node is 1 and all the rest are zero.

Note that the derivative P ′(z) of a matrix polynomial is a straightforward extension to matrices of the ordinary derivative. It is isomorphic to the matrix with entries that are the ordinary derivatives of the original matrix.

The linearization of the previous section changes to the following elegant form. The matrix C1 is unchanged,

0 1   (3.46) C = .. , 1  .     1     1     being (ℓ + 2) by (ℓ + 2) as before. The matrix C0 changes, picking up transposed Jordan-like blocks for each distinct node. For instance, suppose we have two distinct nodes, τ0 and τ1. Suppose further that τ0 has ′ ′′ confluency s0 = 3 while τ1 has confluency s1 = 2. This means that we know f(τ0), f (τ0)/1!, f (τ0)/2!, f(τ1) ′ and f (τ1)/1!. Then,

′′ ′ ′ 0 −f (τ0)/2! −f (τ0)/1! −f(τ0) −f (τ1)/1! −f(τ1) β τ  02 0  β01 1 τ0 (3.47) C0 =   .  β00 1 τ0     β11 τ1     β 1 τ   10 1    Note the reverse ordering of the derivative values in this formulation.

The matrices E(z) and F (z) demonstrating that this is indeed a local linearization have, so far as we know, not been noted in the literature. However, they are exactly the same as those for the Lagrange basis, mutatis mutandis, which are discussed in the appendix to [2],with appropriately modified meanings for φ and D. The new φ contains the Hermite interpolational bases in Equation (3.45), and now D is not diagonal, but rather block diagonal with the transposed Jordan-like blocks above. Both (rational) matrices are still unimodular. Again we have E(z)(zC1 − C0)F (z) = diag(P (z), In,..., In), everywhere except the nodes. 14 To generalize the argument of the previous section that the companion pencil is equivalent also at the nodes requires more work than the Lagrange case does, because the Jordan-like block structure interferes with the perturbation argument; on the other hand, one only needs the highest derivatives at each node to be nonsingular for the construction to work. We leave the details for another paper.

We believe that the strict equivalence proof for the Lagrange basis in [9] can be extended to the Hermite case as well, but this has not yet been carried out.

For the standard triple, take in the scalar case

T (3.48) Y = 1 0 ··· 0

but for X take the coefficients of the expansion of the polynomial 1 in this particular Hermite interpolational basis: it is equal to 1 at each node but has all derivatives zero at each node. That is, put

ρij =1 if j =0 , (3.49) (0 otherwise , and sort them in order:

(3.50) X = 0 ρ0,s0−1 ρ0,s0−2 ··· ρ0,0 ρ1,s1−1 ··· ρn,0 .

For the earlier instance (two nodes, of confluency 3 and 2, respectively), 

(3.51) X = 0 0 0 1 0 1 .

for for Then,  τ0 τ1

−1 | {z } |{z−}1 (3.52) p (z)= X(zC1 − C0) Y .

Remark 3.2. We may re-order the nodes in any fashion we like, and each ordering generates its own linearization (both Hermite and Lagrange). We may also find a linearization where the confluent data is ′ ′′ ordered p(τi), p (τi)/1!, p (τi)/2!, etc., although we have not done so. If there is just one node of confluency ℓ, we recover the standard Frobenius companion (plus two infinite roots):

0 −pℓ−1 −pℓ−2 ··· −p1 −p0 0 1 τ0 1     0 1 τ0 .   (3.53) .. ,  ..  .   0 1 .   1      . ..    . . τ0   1     0 1 τ     0    k ℓ p( ) (τ ) k Here, pk = 0 /k! is the ordinary coefficient in the expansion p(z)= k=0 pk(z −τ0) . The numerical sta- bility of these Hermite interpolational linearization has been studied briefly [26] but much remains unknown. P We confine ourselves in this paper to the study of the standard triple.

To make a linearization for matrix polynomials out of these scalar linearizations, take the Kronecker with In, and insert the appropriate matrix polynomial values and derivative values. 15 Example 3.3 (Matrix polynomial case). Let

τ = [0, 1] and z P (z) P ′(z) −1 0 τ =0 0 −1 1   0 1 1 −1 τ =1 1 1 −1 −1 0     Then, the standard triple is 0 0 −11 0 −11 0 00000000 0 0 10 −1 1 1 −1 00000000     1 0 100 000 00100000      0 1 010 000  00010000 C0 =   C1 =   −10 101 000  00001000      0 −1010 100  00000100      1 0 000 000  00000010      0 1 000 000  00000001         1 0 0 1   0 0   00001010 0 0 X = Y =   . 00000101 0 0     0 0   0 0   0 0     The Hermite interpolating polynomial is

z − 1 −2z2 +3z P (z)= −3z2 +5z − 1 2z2 − 4z +1   and the resolvent form is

−2z2 +4z − 1 −2z2 +3z 4 3 2 4 3 2 X (zC − C )−1 Y = 6z − 21z + 23z − 8z +1 6z − 21z + 23z − 8z +1 . 1 0  −3z2 +5z − 1 −z +1   6z4 − 21z3 + 23z2 − 8z +1 6z4 − 21z3 + 23z2 − 8z +1   Then,

−1 X (zC1 − C0) Y P (z)= I2 , which indicates that the standard triples is correct. Remark 3.4. The modified linearizations of [35] also have standard triples that can be used for algebraic linearization, and arguably should be tabled here as well. They have the advantage of including fewer eigen- values at infinity, or no spurious eigenvalues at infinity, which may lead to better algebraic linearizations. 16 However, they are more involved, and we have less numerical experience with them. In particular, we do not understand their dependence on the ordering of the nodes, and so we leave their analysis to a future study.

4. Concluding remarks. The generalized standard triples (or standard quadruples, if you prefer) that we propose in this paper for convenience in algebraic linearization may have other uses. As pointed out on p. 28 of [20] many of the properties stated in that work for monic polynomials are valid for non- monic polynomials with the appropriate changes made. Some caution with the results of this paper are thus mandated.

We have here defined these generalized standard triples simply by the resolvent representation for the matrix polynomial Equation (1.7), and only for linearizations, which is all we need for algebraic linearization.

The main theorem of the paper, namely Theorem 2.1, gives a universal way to construct this generalized standard triple in any polynomial basis. We also gave explicit instructions for this construction using any of several polynomial bases, for convenience, together with separate proofs using the Schur complement, which may give insight for further work in this area.

In Section 1.3, we have sketched an explicit construction for matrices E(z) and F (z) showing that algebraic linearizations are, in fact, linearizations, with E(z)(zDH − H)F (z) = diag(P (z), I,..., I). We have also given new constructions for matrices E(z) and F (z) which likewise show that the companions for the Lagrange and Hermite interpolational bases are, in fact, linearizations (of matrix polynomials of higher grade). A proof for Lagrange interpolational bases was given already in [2], where indeed the linearization was proved to be strong, but the result for Hermite interpolational bases is new to this paper. We also used Hermite Form computations to give a new (to us) pair E(z) and F (z) for the ordinary monomial basis.

Acknowledgments. We acknowledge the support of Western University, The National Science and Engineering Research Council of Canada, the Ontario Graduate Scholarship (OGS) program, the University of Alcalá, the Ontario Research Centre of Computer Algebra, and the Rotman Institute of Philosophy. Part of this work was developed while RMC was visiting the University of Alcalá, in the frame of the project Giner de los Rios. The authors would also like to thank Peter Lancaster for teaching RMC long ago the value of the 5 × 5 example. Similarly we thank John C. Butcher for the proper usage of the word “interpolational”. We thank Françoise Tisseur for her thorough comments on an earlier version of this paper, and likewise Froilán Dopico for a similarly fruitful discussion. Finally, we thank an anonymous referee for improving (and correcting!) several proofs.

REFERENCES

[1] Maha Al-Ammari and Françoise Tisseur. Standard triples of structured matrix polynomials. and its applications, 437(3):817–834, 2012. [2] Amir Amiraslani, Robert M. Corless, and Peter Lancaster. Linearization of matrix polynomials expressed in polynomial bases. IMA Journal of , 29(1):141–157, 2008. [3] Jean-Paul Berrut and Lloyd N. Trefethen. Barycentric Lagrange interpolation. SIAM Review, 46(3):501–517, 2004. [4] Timo Betcke, Nicholas J. Higham, Volker Mehrmann, Christian Schröder, and Françoise Tisseur. NLEVP: A collection of nonlinear eigenvalue problems. ACM Transactions on Mathematical Software (TOMS), 39(2):7, 2013. [5] Maria Isabel Bueno Cachadina, Javier Perez, Anthony Akshar, Daria Mileeva, and Remy Kassem. Linearizations for inter- polatory bases - a comparison: New families of linearizations. The Electronic Journal of Linear Algebra, 36(36):799– 833, December 2020. [6] J. M. Carnicer, Y. Khiar, and J. M. Peña. Optimal stability of the Lagrange formula and conditioning of the Newton formula. Journal of Approximation Theory, 2017. 17 [7] Eunice Y. S. Chan, Robert M. Corless, Laureano Gonzalez-Vega, J. Rafael Sendra, and Juana Sendra. Algebraic lineariza- tions for matrix polynomials. Linear Algebra and its Applications, 563:373–399, 2019. [8] Robert M. Corless and Nicolas Fillion. Polynomial and rational interpolation. In A Graduate Introduction to Numerical Methods, pages 331–401. Springer, 2013. [9] Robert M. Corless, Leili Rafiee Sevyeri, and B. David Saunders. Equivalences for linearizations of matrix polynomials. https://arxiv.org/abs/2102.09726. [10] Robert M Corless, Nargol Rezvani, and Amirhossein Amiraslani. Pseudospectra of matrix polynomials that are expressed in alternative bases. Mathematics in Computer Science, 1(2):353–374, 2007. [11] Robert M. Corless and Stephen M. Watt. Bernstein bases are optimal, but, sometimes, Lagrange bases are better. In Proceedings of SYNASC, Timisoara, pages 141–153. MIRTON Press, 2004. [12] Froilán M Dopico, Piers W. Lawrence, Javier Pérez, and Paul Van Dooren. Block Kronecker linearizations of matrix polynomials and their backward errors. Numerische Mathematik, 140(2):373–426, 2018. [13] Froilán M. Dopico, Silvia Marcaida, María C. Quintana, and Paul Van Dooren. Local linearizations of rational matrices with application to rational approximations of nonlinear eigenvalue problems. Linear Algebra and its Applications, 604:441–475, November 2020. [14] Rida T. Farouki. The basis: A centennial retrospective. Computer Aided Geometric Design, 29(6):379–419, 2012. [15] Rida T. Farouki and T. Goodman. On the optimal stability of the Bernstein basis. Mathematics of Computation of the American Mathematical Society, 65(216):1553–1566, 1996. [16] Rida T. Farouki and V. T. Rajan. On the numerical condition of polynomials in Bernstein form. Computer Aided Geometric Design, 4(3):191–216, 1987. [17] Heike Faßbender and Philip Saltenberger. On vector spaces of linearizations for matrix polynomials in orthogonal bases. Linear Algebra and its Applications, 525:59–83, July 2017. [18] Walter Gautschi. Orthogonal polynomials in MATLAB: Exercises and Solutions, volume 26. SIAM, 2016. [19] Israel Gohberg, Peter Lancaster, and Leiba Rodman. Indefinite Linear Algebra and Applications. Springer, 2005. [20] Israel Gohberg, Peter Lancaster, and Leiba Rodman. Matrix Polynomials. SIAM Classics in Applied Mathematics, 2009. [21] Stefan Güttel and Françoise Tisseur. The nonlinear eigenvalue problem. Acta Numerica, 26:1–94, 2017. [22] David J. Jeffrey and Robert M. Corless. Linear algebra in Maple. In Leslie Hogben, editor, Handbook of Linear Algebra, chapter 89. Chapman and Hall/CRC, 2013. [23] Guðbjörn F. Jónsson. Eigenvalue methods for accurate solution of polynomial equations. PhD thesis, Center for Applied Mathematics, Cornell University, Ithaca, NY, 2001. [24] Guðbjörn F. Jónsson and Stephen Vavasis. Solving polynomials with small leading coefficients. SIAM Journal on Matrix Analysis and Applications, 26(2):400–414, 2004. [25] V. N. Kublanovskaya. Methods and algorithms of solving spectral problems for polynomial and rational matrices. Journal of Mathematical Sciences, 96(3):3085–3287, 1999. [26] Piers W. Lawrence and Robert M. Corless. Numerical stability of barycentric Hermite root-finding. In Proceedings of the 2011 International Workshop on Symbolic-Numeric Computation, pages 147–148. ACM, 2012. [27] Jörg Liesen and Christian Mehl. Matrix polynomials. In Leslie Hogben, editor, Handbook of Linear Algebra, chapter 18. Chapman and Hall/CRC, 2013. [28] D. Steven Mackey, Niloufer Mackey, Christian Mehl, and Volker Mehrmann. Vector spaces of linearizations for matrix polynomials. SIAM Journal on Matrix Analysis and Applications, 28(4):971–1004, 2006. [29] D. Steven Mackey, Niloufer Mackey, and Françoise Tisseur. Polynomial eigenvalue problems: Theory, computation, and structure. In Numerical Algebra, Matrix Theory, Differential-Algebraic Equations and Control Theory, pages 319–348. Springer, 2015. [30] D. Steven Mackey and Vasilije Perović. Linearizations of matrix polynomials in Bernstein bases. Linear Algebra and its Applications, 501:162–197, 2016. [31] Yuji Nakatsukasa, Vanni Noferini, and Alex Townsend. Vector spaces of linearizations for matrix polynomials: a bivariate polynomial approach. SIAM Journal on Matrix Analysis and Applications, 38(1):1–29, 2017. [32] Leonardo Robol, Raf Vandebril, and Paul Van Dooren. A framework for structured linearizations of matrix polynomials in various bases. SIAM Journal on Matrix Analysis and Applications, 38(1):188–216, 2017. [33] Arne Storjohann. An O(n3) algorithm for the . In Proceedings of the 1998 international symposium on Symbolic and algebraic computation, pages 101–105. ACM, 1998. [34] Olga Taussky and Hans Zassenhaus. On the similarity transformation between a matrix and its transpose. Pacific Journal of Mathematics, 9(3):893–896, 1959. [35] Roel Van Beeumen, Wim Michiels, and Karl Meerbergen. Linearization of Lagrange and Hermite interpolating matrix polynomials. IMA Journal of Numerical Analysis, 35(2):909–930, 2015.

18